• Nenhum resultado encontrado

Search for a very light NMSSM Higgs boson produced in decays of the 125 GeV scalar boson and decaying into $\tau$ leptons in pp collisions at $\sqrt{s}=8$ TeV

N/A
N/A
Protected

Academic year: 2021

Share "Search for a very light NMSSM Higgs boson produced in decays of the 125 GeV scalar boson and decaying into $\tau$ leptons in pp collisions at $\sqrt{s}=8$ TeV"

Copied!
44
0
0

Texto

(1)

EUROPEAN ORGANIZATION FOR NUCLEAR RESEARCH (CERN)

CERN-PH-EP/2013-037 2016/01/21

CMS-HIG-14-019

Search for a very light NMSSM Higgs boson produced in

decays of the 125 GeV scalar boson and decaying into

τ

leptons in pp collisions at

s

=

8 TeV

The CMS Collaboration

Abstract

A search for a very light Higgs boson decaying into a pair of τ leptons is presented within the framework of the next-to-minimal supersymmetric standard model. This search is based on a data set corresponding to an integrated luminosity of 19.7 fb−1of proton-proton collisions collected by the CMS experiment at a centre-of-mass energy of 8 TeV. The signal is defined by the production of either of the two lightest scalars, h1or h2, via gluon-gluon fusion and subsequent decay into a pair of the lightest Higgs

bosons, a1or h1. The h1or h2boson is identified with the observed state at a mass of

125 GeV. The analysis searches for decays of the a1(h1)states into pairs of τ leptons

and covers a mass range for the a1 (h1)boson of 4 to 8 GeV. The search reveals no

significant excess in data above standard model background expectations, and an upper limit is set on the signal production cross section times branching fraction as a function of the a1(h1)boson mass. The 95% confidence level limit ranges from 4.5 pb

at ma1 (mh1) =8 GeV to 10.3 pb at ma1 (mh1) =5 GeV.

Published in the Journal of High Energy Physics as doi:10.1007/JHEP01(2016)079.

c

2016 CERN for the benefit of the CMS Collaboration. CC-BY-3.0 license

See Appendix A for the list of collaboration members

(2)
(3)

1

1

Introduction

The recently discovered particle with mass close to 125 GeV [1–3] has been shown to have properties that are consistent with those of a standard model (SM) Higgs boson [4–14]. Su-persymmetric (SUSY) extensions of the SM [15, 16] also predict a particle with such proper-ties and resolve some problems of the SM [17]. The minimal supersymmetric standard model (MSSM) [18, 19] postulates the existence of two Higgs doublets, resulting in five physical states: two CP-even, one CP-odd, and two charged Higgs bosons. This version of SUSY has been ex-tensively tested using data collected by the ATLAS and CMS experiments at the CERN LHC. However, nonminimal SUSY extensions have received far less attention. One example is the next-to-MSSM (NMSSM), which extends the MSSM by an additional singlet superfield, inter-acting only with itself and the two Higgs doublets [18, 20–26]. This scenario has all the desirable features of SUSY, including a solution of the hierarchy problem and gauge coupling unification. In the NMSSM, the Higgs mixing parameter µ is naturally generated at the electroweak scale through the vacuum expectation value of the singlet field, thereby solving the so-called µ prob-lem of the MSSM [27]. Furthermore, the amount of fine tuning required in the NMSSM to obtain a CP-even Higgs boson with a mass of 125 GeV is significantly reduced compared to the MSSM [28–30]. The Higgs sector of the NMSSM is larger than that of the MSSM. There are seven Higgs bosons: three CP-even(h1,2,3), two CP-odd(a1,2), and two charged Higgs states.

By definition, mh3 > mh2 > mh1 and ma2 > ma1. Over large parts of the NMSSM

parame-ter space, the observed boson with mass close to 125 GeV, hereafparame-ter denoted H(125), could be identified with one of the two lightest scalar NMSSM Higgs bosons, h1or h2.

A vast set of next-to-minimal supersymmetric models is consistent with the SM measurements and constraints from searches for SUSY particles made with LHC, Tevatron, SLAC and LEP data, as well as with the properties of the H(125)boson measured using Run 1 LHC data [31– 36]. These models provide possible signatures that cannot be realized in the MSSM given recent experimental constraints [37]. For example, the decays H(125) → h1h1 and H(125) → a1a1

are allowed when kinematically possible. These decay signatures have been investigated in phenomenological studies considering a variety of production modes at the LHC [38–45]. The analysis presented in this paper is motivated by the NMSSM scenarios that predict a very light h1or a1 state with mass in the range 2mτ < mh1 (ma1) < 2mb, where mτ is the mass of the τ lepton and mb is the mass of the b quark. Such a light state is potentially accessible in final states with four τ leptons, where H(125) → h1h1(a1a1) → 4τ [46, 47]. In these scenarios the

decay H(125) →a2a2is not kinematically allowed.

Several searches for H(125) → φ1φ1 decays, where φ1 can be either the lightest CP-even state

h1or the lightest CP-odd state a1, have been performed. The analyses carried out by the OPAL

and ALEPH Collaborations at LEP [48, 49] searched for the decay of the CP-even Higgs boson into a pair of light CP-odd Higgs bosons, exploiting the Higgs-strahlung process, where the CP-even state is produced in association with a Z boson. These searches found no evidence for a signal, and limits were placed on the signal production cross section times branching fraction. However, searches at LEP did not probe masses of the CP-even state above 114 GeV. A similar study has been performed by the D0 Collaboration at the Tevatron [50], searching for inclusive production of the CP-even Higgs boson in pp collisions followed by its decay into a pair of light CP-odd Higgs bosons. No signal was detected and upper limits were set on the signal production cross section times branching fraction in the mass ranges 3.6 < ma1 < 19 GeV and

89 < mH < 200 GeV. The limits set by the D0 analysis are a factor one to seven times higher

compared to the SM production cross section for pp→H(125) +X.

(4)

2 2 Signal topology

in decays of a heavier CP-even state [51]. This study probed the mass of the CP-odd state in the range 2mµ<ma1 <2mτ, where mµis the mass of muon. In this mass range the decay a1 →µµ can be significant. No evidence for a signal was found and upper limits were placed on the signal production cross section times branching fraction. The ATLAS Collaboration has also recently searched for h/H→a1a1→µµττ[52], covering the mass range ma1 =3.7–50 GeV for

mH =125 GeV, and mH=100–500 GeV for ma1 =5 GeV. No excess over SM backgrounds was

observed, and upper limits were placed on σ(gg → H) B(H → a1a1) B2(a1 → ττ), under the

assumption that Γ(a→µµ) Γ(a→ττ) = m 2 µ m2 τ q 1− 2mτ/ma 2 .

The search for the production of a pair of light bosons with their subsequent decay into four τ leptons has not yet been performed at the LHC and is the subject of this paper. The choice of the 4τ channel makes it possible to probe the signal cross section times branching fraction

(σB)sig≡σ(gg→H(125)) B(H(125) →φ1φ1) B2(φ1→ττ)

in a model-independent way.

2

Signal topology

This paper describes a search for the production of the H(125)boson, with its decay into a pair of light NMSSM Higgs bosons φ1. The signal can be associated with one of three possible

scenarios:

• H(125)corresponds to h2and decays into a pair of h1states, h2→h1h1; • H(125)corresponds to h2and decays into a pair of a1states, h2 →a1a1; • H(125)corresponds to h1and decays into a pair of a1states, h1 →a1a1.

The analysis is optimized for the gluon-gluon fusion process, which is the dominant production mechanism of the H(125)boson at the LHC. The signal topology is illustrated in Fig. 1. The search is performed for very light φ1 states, covering a mass range of 4 to 8 GeV. Within this

mass range the φ1boson is expected to decay predominantly into a pair of τ leptons, φ1 →ττ.

In the decay of each φ1, one of the τ leptons is identified via its muonic decay. The other

τ lepton is required to decay into a one-prong mode, i.e. a decay into one charged particle

(electron, muon, or hadron) and one or more neutral particles. We identify these decays by the presence of one reconstructed track with charge sign opposite to that of the closest muon. Neutral particles are not considered in the event selection.

Given the large difference in mass between the φ1and the H(125)states (mH(125)  mφ1), one

expects the φ1 bosons to have large Lorentz boosts and their decay products to be collimated.

Furthermore, in the gluon-gluon fusion process the H(125)state is mainly produced with rela-tively small transverse momentum pT. Thus, in the majority of H(125) → φ1φ1decays, the φ1

states would be produced nearly back-to-back in the plane transverse to the beams, with a large separation in azimuthal angle φ between the decay products of the two φ1bosons. The H(125)

state can be produced with relatively high transverse momentum if a hard gluon is radiated from the initial-state gluons or the heavy-quark loop. In this case, the separation between two

(5)

3

φ1bosons in azimuthal angle is reduced, while the separation in pseudorapidity η can still be

large. The pseudorapidity is defined via the polar angle θ as η≡ −ln[tan(θ/2)].

The φ1 → ττdecays into final states without muons are not considered in the analysis. These

decays are mimicked by hadronic jets with a significantly higher probability compared to final states with at least one muon and contribute marginally to the search sensitivity.

The signal properties discussed above are used to define the search topology. The analysis pre-sented here searches for the signal in a sample of dimuon events with large angular separation between the muons. The two muons are required to have the same sign. This criterion almost entirely eliminates background from the Drell–Yan process, gauge boson pair production, and tt production. Each muon is accompanied by one nearby opposite-sign track. Further details of the kinematic selection are given in Section 4. Throughout this paper, the signal yields are normalized to the benchmark value of the signal production cross section times branching frac-tion of 5 pb. The choice of the benchmark scenario is motivated by recent phenomenological analyses [46, 47].

Lorentz-boosted states

well separated same-sign muons

Figure 1: Left: Feynman diagram for the signal process. Right: Illustration of the signal topol-ogy. The label “µ∓/e∓/h∓” denotes a muon, electron, or charged-hadron track.

3

CMS detector, data, and simulated samples

The central feature of the CMS apparatus is a superconducting solenoid of 6 m internal diam-eter, providing a field of 3.8 T. The innermost component of the detector is a silicon pixel and strip tracker, which is used to measure the momenta of charged particles and reconstruct col-lision vertices. The tracker, which covers the pseudorapidity range|η| < 2.5, is surrounded

by a crystal electromagnetic calorimeter and a brass and scintillator hadronic calorimeter, both placed inside the solenoid. These calorimeters cover|η| < 3.0. A quartz fiber Cherenkov

for-ward hadron detector extends the calorimetric coverage to|η| < 5.0. Muons are measured in

the pseudorapidity range|η| <2.4, with detection planes made using three technologies: drift

tubes, cathode strip chambers, and resistive-plate chambers. The first level of the CMS trigger system, composed of custom hardware processors, uses information from the calorimeters and muon detectors to select the most interesting events in a fixed time interval of 4 µs. The high-level trigger processor farm further decreases the event rate from around 100 kHz to less than 1 kHz before data storage. A more detailed description of the CMS detector, together with a definition of the coordinate system used and the relevant kinematic variables, can be found in Ref. [53].

(6)

4 4 Event selection

The data set used in this analysis was recorded in 2012 and corresponds to an integrated lumi-nosity of 19.7 fb−1of pp collisions at√s=8 TeV.

The Monte Carlo (MC) event generatorPYTHIA6.426 [54] is used to model the NMSSM Higgs boson signal produced via gluon-gluon fusion. The H(125)boson pTspectrum fromPYTHIAis

reweighted to the spectrum obtained from a leading-order computation with a next-to-next-to-leading logarithmic accuracy using the HQT 2.0 program [55, 56], which performs the resummation of the large logarithmic contributions appearing at transverse momenta much smaller than the mass of the Higgs boson. For optimisation studies, diboson and quantum chromodynamics (QCD) multijet backgrounds are simulated by PYTHIA. Inclusive Z, W, and tt production are modelled with MADGRAPH5.1 [57]. The MADGRAPHgenerator is interfaced with PYTHIAfor parton showering and fragmentation. The PYTHIA parameters that steer the simulation of hadronisation and the underlying event are set to the most recent PYTHIA Z2* tune. This tune is derived from the Z1 tune [58], which uses the CTEQ5L parton distribution function (PDF) set, whereas Z2* adopts the CTEQ6L PDF set [59]. TheTAUOLApackage [60] is used for τ lepton decays in all cases. All generated events, with the exception of a few special QCD multijet samples discussed in Section 6.2, are processed through a detailed simulation of the CMS detector, based on GEANT4 [61], and are reconstructed employing the same algo-rithms as for data.

4

Event selection

Events are recorded using double-muon triggers with thresholds on the muon transverse mo-menta of 17 GeV for the leading muon and 8 GeV for the subleading one. To pass the high-level trigger, the tracks of the two muons are additionally required to have points of closest approach to the beam axis within 2 mm of each other along the longitudinal direction.

In 2012, the average number of pp interactions per LHC bunch crossing (pileup) was about 20. The simulated MC events are reweighted to represent the distribution of the number of pileup interactions per bunch crossing in data.

For each reconstructed collision vertex, the sum of the pT2 of all tracks associated with the

vertex is computed. The vertex for which this quantity is largest is assumed to correspond to the hard-scattering process, and is referred to as the primary vertex (PV).

The identification and reconstruction of muons is achieved by matching track segments found in the silicon tracker with those found in the muon detectors [62]. Additional requirements are applied on the number of measurements in the inner pixel and outer silicon strip detectors, on the number of matched segments in the muon detectors, and on the quality of the global muon track fit, quantified by χ2.

The data are further selected by requiring at least one pair of muons with the same charge. This requirement significantly suppresses background contributions originating from the Drell–Yan process, from decays of tt pairs, and from QCD multijet events with muonic decays of heavy-flavour hadrons. The leading muon is required to have pT > 17 GeV and |η| < 2.1. The

subleading muon is required to have pT >10 GeV and|η| <2.1. To reject QCD multijet events

with muonic decays of hadrons containing charm or bottom quarks, selections are applied on the impact parameters of the muon tracks. The impact parameter in the transverse plane is required to be smaller than 300 µm with respect to the PV. The longitudinal impact parameter is required to be smaller than 1 mm with respect to the PV. The two selected same-sign muons are required to be separated by∆R(µ, µ) =

(7)

5

pseudorapidity and∆φ is the separation in azimuthal angle between the two muons. If more than one same-sign muon pair is found in the event, the pair with the largest scalar sum of muon transverse momenta is chosen.

The analysis makes use of reconstructed tracks that fulfill selection criteria based on the track fit quality, the number of measurements in the inner pixel and outer strip silicon detector, and track impact parameters with respect to the PV [63]. Tracks must have pT > 1 GeV and|η| <

2.4. The impact parameter in the transverse plane and the longitudinal impact parameter are required to be smaller than 1 cm relative to the PV.

Given the search topology, we require each muon to be accompanied by exactly one track sat-isfying these criteria within a ∆R cone of radius 0.5 centred on the muon direction. We label such muon-track pairs as “isolated”.

The loose impact parameter requirements on the tracks are designed to suppress background events in which a heavy-flavour hadron decays into a muon and several charged particles. Although tracks from these decay products will be displaced from the PV, they can still satisfy the loose track impact parameter criteria. Such events are rejected by the requirement of exactly one track accompanying the muon.

The track around each muon is identified as a one-prong τ lepton decay candidate if it fulfils the following selection criteria.

• The nearby track is required to have charge opposite to the muon.

• The track must have pT >2.5 GeV and|η| <2.4.

• The transverse and longitudinal impact parameters of the track are required to be smaller than 200 µm and 400 µm relative to the PV, respectively.

5

Signal extraction

The set of selection requirements outlined in the previous section defines the signal region. The number of selected data events, the expected background and signal yields, and the signal ac-ceptances after selection in the signal region are reported in Table 1. The expected background and signal yields, along with the signal acceptances, are obtained from simulation. The sig-nal yields are normalized to the benchmark value of the sigsig-nal production cross section times branching fraction of 5 pb. The quoted uncertainties in predictions from simulation include only MC statistical uncertainties. It should be noted that no MC simulation is used to evaluate the background in the analysis described below as the modelling is based fully on data. The ex-pected background yields presented in Table 1 show that the final selected sample is dominated by QCD multijet events, and that the contribution from other background sources is negligible, constituting less than 1% of all selected events. Although MC simulation is not directly used to estimate background, the simulated samples play an important role in the validation of the background modelling as described in Section 6. The signal acceptances are computed with respect to all possible decays of the four τ leptons, and include a branching fraction factor

1 2B

2(

φ1→τµτone-prong) ≈3.5%,

where the factor 1/2 accounts for the selection of same-sign muon pairs, andB(φ1→τµτone-prong) denotes the branching fraction of the φ1 → ττ decays to the final states characterized by the

(8)

6 5 Signal extraction

Table 1: The number of observed events, expected background and signal yields, and signal ac-ceptances after final selection. The computed signal acac-ceptances include the branching fraction factorB2(

φ1 → τµτone-prong)/2. The electroweak background contribution includes the Drell–

Yan process, W+jets production, and diboson production of WW, WZ, and ZZ. The numbers of signal events are reported for the benchmark value of the signal production cross section times branching fraction of 5 pb. The expected background and signal yields and signal accep-tances are obtained from simulation. The quoted uncertainties in predictions from simulation include only statistical uncertainties related to the size of MC samples.

Sample Signal acceptance Number of events

A(gg→H(125) →φ1φ1→)

Signal for(σB)sig=5 pb

mφ1 =4 GeV (5.38±0.23) ×10 −4 53.0±2.31 =5 GeV (4.36±0.21) ×10−4 43.0±2.0 mφ1 =6 GeV (4.00±0.23) ×10−4 39.5±2.0 mφ1 =7 GeV (4.04±0.20) ×10 −4 39.9±2.01 =8 GeV (3.13±0.18) ×10−4 30.8±1.8 Background QCD multijet — 820±320 tt — 1.2±0.2 Electroweak — 5.0±4.7 Data — 873

This branching fraction is expressed as

B(φ1 →τµτone-prong) =2B(τ→one-prong)B(τµνν) − B2(τµνν),

whereB(τ → one-prong)denotes the total branching fraction of the τ decay to one charged

particle with any number of neutral particles. The factor of two in the first term accounts for the two possible charges of the required muonic decay: ττ+ → µ−+one-prong+ and ττ+ → µ++one-prong−. Subtraction of the termB2(τµνν)avoids double counting in

the case where the two τ leptons produced by a given φ1both decay to muons.

The invariant mass of each selected muon and the nearby track is reconstructed. The two-dimensional distribution of the invariant mass of each selected muon and the nearby track is used to discriminate between the signal and the QCD multijet background; the signal is extracted by means of a fit to this dimensional distribution. The binning of the two-dimensional (m1, m2) distributions is illustrated in Fig. 2. For masses below 3 GeV, bins of

1 GeV width are used for both m1 and m2. For masses in the range 3 < m1(m2) < 10 GeV, a

single bin is used. This choice avoids poorly populated bins in the two-dimensional (m1, m2)

distributions in the background control regions used to construct and validate the QCD multi-jet background model (Section 6). For each selected event, the (m1, m2) histogram is filled once

if the pair of quantities (m1, m2) occurs in one of the diagonal bins and twice, once with

val-ues (m1, m2) and a second time with the swapped values (m2, m1), for off-diagonal bins. This

procedure insures the symmetry of the two-dimensional (m1, m2) distribution. To avoid double

counting of events, the off-diagonal bins(i, j)with i> j are excluded from the procedure of the signal extraction (the hatched bins in Fig. 2). Thus, the number of independent bins is reduced

(9)

7 from 4×4=16 to 4× (4+1)/2=10.

(GeV)

1

m

0

1

2

3

4

5

6

7

8

9

10

(GeV)

2

m

0

1

2

3

4

5

6

7

8

9

10

(1,1) (1,2) (1,3) (1,4) (2,2) (2,3) (2,4) (3,3) (3,4) (4,4)

Figure 2: Binning of the two-dimensional (m1, m2) distribution. The hatched bins are excluded

from the statistical analysis, as detailed in the text.

In order to fit the data in the 10 bins of the dimensional distribution of Fig. 2, a two-component fit is performed using two-dimensional distributions (“templates”) describing the QCD multijet background and the signal. The normalisations of background and signal com-ponents are free parameters in this fit. The two-dimensional template for the signal is obtained from the simulation using the generator described in Section 3. The two-dimensional template for the QCD multijet background is extracted from the data as explained in the next section.

6

Modelling of the QCD multijet background shape

A simulation study shows that the sample of same-sign dimuon events selected as described in Section 4, but without requiring a presence of one-prong τ candidates and without apply-ing the isolation requirement for the muon-track systems, is dominated by QCD multiparton production, where 94% of all selected events contain b quarks in the final state. The same-sign muon pairs in these events originate mainly in the following cases.

• Muonic decay of a bottom hadron in one b quark jet, and cascade decay of a bottom hadron into a charmed hadron with subsequent muonic decay of a charmed hadron in the other b quark jet.

• Muonic decay of a bottom hadron in one b quark jet, and decay of a quarkonium state into a pair of muons in the other jet.

• Muonic decay of a bottom hadron in one b quark jet, and muonic decay of a neutral B meson in the other b quark jet. The same-sign muon pair in this case may appear as a result of B0–B0oscillations.

(10)

8 6 Modelling of the QCD multijet background shape

The normalization of the QCD multijet background is not constrained prior to the extraction of the signal. The procedure used to model the shape of the two-dimensional (m1, m2) distribution

of QCD multijet events in the signal region is described in this section.

Given the symmetry of the two-dimensional (m1, m2) distribution, the modelling of the QCD

multijet background shape is derived from the two-dimensional probability density function (pdf)

f2D(m1, m2) =C(m1, m2)f1D(m1)f1D(m2), (1)

where

• f2D(m1, m2)is the two-dimensional pdf of the invariant masses of the muon-track

systems, m1 and m2, in the sample of QCD multijet events selected in the signal

region;

• f1D(mi)is the one-dimensional pdf of the invariant mass of the muon-track system

in the sample of QCD multijet events selected in the signal region;

• C(m1, m2)is a symmetric function of two arguments, C(m1, m2) = C(m2, m1),

re-flecting the correlation between m1and m2.

A constant correlation function would indicate the absence of correlation between m1and m2.

Based on Eq. (1), the content of bin(i, j)of the symmetric normalized two-dimensional distri-bution f2D(m1, m2)is computed as

f2D(i, j) =C(i, j)f1D(i)f1D(j), (2)

where

• C(i, j)is the correlation coefficient in the bin(i, j)of the correlation function C(m1, m2); • f1D(i)is the content of bin i in the normalized one-dimensional distribution f1D(m).

The modelling of f1D(m)and C(m1, m2), described in the following, is necessary in order to

build the template f2D(i, j). 6.1 Modelling of f1D

(

m

)

The f1D(m)pdf is modelled using a QCD-enriched control data sample disjoint from the signal

region. Events in the control sample are required to satisfy all selection criteria, except for the isolation of the second muon-track system. The second muon is required to be accompanied by either two or three nearby tracks with pT > 1 GeV and impact parameters smaller than

1 cm relative to the PV both in the transverse plane and along the beam axis. The simulation shows that more than 99% of events selected in this control region, hereafter referred to as

N23, are QCD multijet events. The modelling of the f1D(m)pdf is based on the assumption

that the kinematic distributions for the first muon-track system are not affected by the isolation requirement imposed on the second, and therefore the f1D(m)pdf of the isolated muon-track

system is the same in the signal region and the region N23.

A direct test of this assumption, given the limited size of the simulated sample of QCD multijet events, is not conclusive, and a test is therefore performed with an additional control sample. Events are selected in this control sample if one of the muons has at least one track passing the one-prong τ decay candidate criteria within a ∆R cone of radius 0.5 around the muon

(11)

6.2 Modelling ofC(m1, m2) 9

direction, with any number of additional tracks within the same∆R cone. As more than one of these tracks can pass the selection criteria for a one-prong τ decay candidate, we investigate two scenarios. In one scenario, the lowest pT (“softest”) track passing the one-prong τ decay

candidate criteria is used to calculate the muon-track invariant mass, while in the other scenario the highest pT (“hardest”) track passing the one-prong τ decay candidate criteria is used. If

only one τ track is found around the first muon, the track is regarded as both “hardest” and “softest”. For the second muon, two isolation requirements are considered: when the muon is accompanied by only one track passing the one-prong τ decay candidate criteria (Ntrk,2 =

1) as in the signal region, or when it is accompanied by two or three tracks (Ntrk,2 = 2, 3)

with pT > 1 GeV and impact parameters smaller than 1 cm relative to the PV as in the region N23. The shapes of invariant mass distributions of the first muon and the softest or hardest

accompanying track are then compared for the two different isolation requirements on the second muon, Ntrk,2 = 1 and Ntrk,2 = 2, 3. The test is performed both on data and on the simulated sample of QCD multijet events. The results of this study are illustrated in Fig. 3. In all considered cases, the shape of the invariant mass distribution is compatible within statistical uncertainties between the two cases, Ntrk,2 =1 and Ntrk,2 =2, 3. This observation validates the

assumption that the f1D(m)pdf can be determined in the control region N23.

Figure 4 presents the normalized invariant mass distribution of the muon-track system for data selected in the signal region, and for the QCD multijet background model derived from the control region N23. The data and QCD multijet background distributions are compared

to the signal distribution normalized to unity (signal pdf), obtained from simulation, for two representative mass hypotheses, mφ1 = 4 and 8 GeV. The invariant mass of the muon-track system is found to have high discrimination power between the QCD multijet background and signal at mφ1 =8 GeV. At smaller mφ1the signal shape becomes more similar to the background shape, resulting in a reduction of discrimination power. The normalized distribution f1D(i)

with the binning defined in Fig. 2 is extracted from the background distribution shown in Fig. 4.

6.2 Modelling ofC

(

m1, m2

)

In order to determine the correlation coefficients C(i, j)we define an additional control region

A enriched in QCD multijet events. This control region consists of events that contain two same-sign muons passing the identification and kinematic selection criteria outlined in Sec-tion 4. Each muon is required to have two or three nearby tracks within a∆R cone of radius 0.5 around the muon direction. One and only one of these tracks must satisfy the criteria im-posed on one-prong τ lepton decay candidates with pT >2.5 GeV. The additional tracks must

have transverse momentum in the range 1 < pT < 2.5 GeV. A total of 9127 data events are

selected in this control region. The MC simulation predicts that the QCD multijet background dominates in region A, comprising more than 99% of all selected events. The simulation study also shows that the overall background-to-signal ratio is enhanced compared to the signal re-gion by a factor of 15 to 20, depending on the mass hypothesis mφ1. Despite the large increase in the overall background-to-signal ratio, potential signal contamination in individual bins of the mass distributions can be nonnegligible. Bin-by-bin signal contamination in region A is dis-cussed in Section 7. For each event in control region A, the pair (m1,m2) of muon-track invariant

masses is calculated. This pair is used to build the symmetrized normalized two-dimensional distribution f2D(i, j)defined in Fig. 2. Then C(i, j)is obtained according to Eq. (2) as

C(i, j) = f2D(i, j)

f1D(i)f1D(j)

(12)

10 6 Modelling of the QCD multijet background shape (GeV) 1, softest track µ m ) -1 dN/dm (GeV × 1/N 0.0 0.1 0.2 0.3 0.4 0.5 Data = 1 trk,2 N = 2,3 trk,2 N (8 TeV) -1 19.7 fb CMS (GeV) 1, softest track µ m 0 2 4 6 8 10 ratio 0.5 1.0 1.5 (GeV) 1, hardest track µ m ) -1 dN/dm (GeV × 1/N 0.0 0.1 0.2 0.3 0.4 0.5 Data = 1 trk,2 N = 2,3 trk,2 N (8 TeV) -1 19.7 fb CMS (GeV) 1, hardest track µ m 0 2 4 6 8 10 ratio 0.5 1.0 1.5 (GeV) 1, softest track µ m ) -1 dN/dm (GeV × 1/N 0.0 0.1 0.2 0.3 0.4 0.5 QCD multijets = 1 trk,2 N = 2,3 trk,2 N 8 TeV CMS Simulation (GeV) 1, softest track µ m 0 2 4 6 8 10 ratio 0.5 1.0 1.5 (GeV) 1, hardest track µ m ) -1 dN/dm (GeV × 1/N 0.0 0.1 0.2 0.3 0.4 0.5 QCD multijets = 1 trk,2 N = 2,3 trk,2 N 8 TeV CMS Simulation (GeV) 1, hardest track µ m 0 2 4 6 8 10 ratio 0.5 1.0 1.5

Figure 3: Normalized invariant mass distributions of the first muon and the softest (left plots) or hardest (right plots) accompanying track for different isolation requirements imposed on the second muon: when the second muon has only one accompanying track (Ntrk,2 = 1; squares);

or when the second muon has two or three accompanying tracks (Ntrk,2=2, 3; circles). The

up-per plots show distributions obtained from data. The lower plots show distributions obtained from the sample of QCD multijet events generated with PYTHIA. Lower panels in each plot show the ratio of the Ntrk,2 =1 distribution to the Ntrk,2=2, 3 distribution.

(13)

6.2 Modelling ofC(m1, m2) 11

(GeV)

, track µ

m

)

-1

1/N x dN/dm (GeV

0.0

0.1

0.2

0.3

0.4

0.5

0.6

(8 TeV) -1 19.7 fb

CMS

Data

QCD bkg model

) = 4 GeV

1

φ

Signal, m(

) = 8 GeV

1

φ

Signal, m(

(GeV)

, track µ

m

0

2

4

6

8

10

data/bkg

0.0 0.5 1.0 1.5 2.0

Figure 4: Normalized invariant mass distribution of the muon-track system for events passing the signal selection. Data are represented by points. The QCD multijet background model is derived from the control region N23. Also shown are the normalized distributions from signal

simulations for two mass hypotheses, mφ1 = 4 GeV (dotted histogram) and 8 GeV (dashed histogram). Each event contributes two entries to the distribution, corresponding to the two muon-track systems passing the selection requirements. The lower panel shows the ratio of the distribution observed in data to the distribution, describing the background model.

(14)

12 6 Modelling of the QCD multijet background shape

where f1D(i) is the one-dimensional normalized distribution with two entries per event (m1

and m2) built as for Fig. 4. Correlation coefficients C(i, j)derived from data in region A are

presented in Fig. 5. 0.03 ± 0.98 0.02 ± 0.99 0.03 ± 1.05 0.05 ± 1.06 0.02 ± 0.99 0.03 ± 1.03 0.03 ± 0.97 0.04 ± 0.98 0.03 ± 1.05 0.03 ± 0.97 0.07 ± 1 0.07 ± 0.91 0.05 ± 1.06 0.04 ± 0.98 0.07 ± 0.91 0.14 ± 1

(GeV)

1

m

0

1

2

3

4

5

6

7

8

9

10

(GeV)

2

m

0

1

2

3

4

5

6

7

8

9

10

(8 TeV) -1 19.7 fb

CMS

Region A

Figure 5: The (m1, m2) correlation coefficients C(i, j) along with their statistical uncertainties,

derived from data in the control region A.

A direct comparison of C(i, j) between the signal region and region A would be impossible in the simulated sample of QCD multijet events because of the very small numbers of events selected in the signal region and in region A. In order to assess the difference in C(i, j)between the signal region and region A, a dedicated MC study is performed, making use of a large ex-clusive sample generated withPYTHIA. The simulation includes only two leading-order QCD multijet production mechanisms: the creation of a bb quark pair via gg → bb and qq → bb. The detector simulation and event reconstruction are not performed for this sample, and the comparison of C(i, j) between the signal region and region A is made using generator-level quantities.

These simplifications are validated by performing a set of consistency tests, making use of the available MC sample of QCD multijet events processed through the full detector simulation and event reconstruction. These tests are performed in a control region B, where each muon is required to have at least one track passing the one-prong τ decay candidate selection cri-teria, i.e. with pT > 2.5 GeV and impact parameters smaller than 200 µm and 400 µm in the

transverse plane and along beam axis, respectively. Along with this requirement each muon is allowed to have one or more tracks within a∆R cone of radius 0.5 around the muon direction, with pT > 1 GeV and impact parameters smaller than 1 cm. Control region B is characterized

by a significantly larger yield of QCD multijet events compared to the signal region and con-trol region A, thus making it possible to perform reliable MC consistency tests and assess the uncertainties in C(i, j). Two scenarios are investigated: 1) muons are paired with the softest one-prong τ decay candidate and 2) muons are paired with the hardest one-prong τ decay candidate. If only one one-prong τ decay candidate is found around a muon, it is regarded

(15)

13

as both “softest” and “hardest”. In both scenarios the correlation coefficients computed using the reconstructed four-momenta of muons and tracks are found to be compatible with those computed using generator-level four-momenta, within statistical uncertainties. Furthermore, the correlation coefficients computed with the inclusive QCD multijet sample are found to be compatible with those computed in the exclusive MC sample including only the gg(qq) →bb production mechanisms. This observation validates the use of the generator-level information and the exclusive bb MC sample to compare C(i, j)between the signal region and control re-gion A. This comparison is presented in Fig. 6. The uncertainties in C(i, j)represent a quadratic sum of the systematic and MC statistical uncertainties. The systematic uncertainties are derived from the control region B. They take into account 1) any differences in C(i, j)calculated using the inclusive QCD multijet sample compared with the exclusive bb sample and 2) any differ-ences in C(i, j) calculated using full detector simulation and event reconstruction compared with the study using generator-level quantities. Within their uncertainties the correlation co-efficients C(i, j)in the signal region and in region A are compatible. We therefore use C(i, j)

derived from data in region A to predict the QCD multijet background shape in the signal region according to Eq. (2).

Bins (1,1) (1,2) (1,3) (1,4) (2,2) (2,3) (2,4) (3,3) (3,4) (4,4)

)

i,j(

C

0.6 0.8 1 1.2 1.4 QCD multijets (PYTHIA) Region A Signal region

Generator study (8 TeV)

Figure 6: The (m1, m2) correlation coefficients C(i, j)determined in the control region A (circles)

and in the signal region (squares) from the MC study carried out at generator level with the exclusive MC sample of QCD multijet events resulting from gg(qq) → bb production mech-anisms. The bin notation follows the definition presented in Fig. 2. The vertical bars include both statistical and systematic uncertainties.

7

Systematic uncertainties

The analysis is affected by various systematic uncertainties, which are classified into two groups. The first group consists of uncertainties related to the background, while the second group includes uncertainties related to the signal. The systematic uncertainties are summarized in Table 2.

(16)

14 7 Systematic uncertainties

7.1 Uncertainties related to background

The estimation of the QCD multijet background is based solely on data and is therefore not affected by imperfections in the simulation of the detector response and inaccuracies in the modelling of the muon and track reconstruction.

The shape of the background in the two-dimensional (m1, m2) distribution is modelled

accord-ing to Eq. (2). The uncertainty in the two-dimensional shape f2D(m1, m2) is dominated by

uncertainties in the correlation coefficients C(i, j) derived in the QCD multijet background-enriched control region A as described in Section 6. The statistical uncertainties in C(i, j)in region A range from 2 to 14%, as seen in Fig. 5. These uncertainties are accounted for in the signal extraction procedure by 10 independent nuisance parameters, one nuisance parameter per bin in the (m1, m2) distribution. The systematic uncertainties related to the extrapolation

of C(i, j) from the control region A to the signal region are derived from the dedicated MC study. The correlation coefficients are found to be compatible between the signal region and the control region A within uncertainties ranging from 2 to 22% (Fig. 6). These uncertainties are accounted for by 10 additional independent nuisance parameters.

The possible contamination of control region A by the signal may bias the estimation of the correlation coefficients and consequently have an impact on the evaluation of the QCD multijet background. The effect is estimated with a conservative assumption on the branching fraction

B(H(125) →φ1φ1) B2(φ1 → ττ)of 32%, which corresponds to the 95% confidence level (CL)

upper limit set by CMS on the branching fraction of the H(125)boson decays to non-standard model particles [4], while the cross section for gluon-gluon fusion is set to the value predicted Table 2: Systematic uncertainties and their effect on the estimates of the QCD multijet back-ground and signal. The effect of the uncertainties in C(i, j)on the total background yield is absorbed by the overall background normalization, which is allowed to vary freely in the fit.

Source Value Affected Type Effect on the

sample total yield

Statistical 2–14% bkg. bin-by-bin — uncertainties in C(i, j) Extrapolation 2–22% bkg. bin-by-bin — uncertainties in C(i, j)

Integrated luminosity 2.6% signal norm. 2.6%

Muon ID and trigger

2% per muon signal norm. 4%

efficiency

Track selection and

5% per track signal norm. 10%

isolation efficiency MC statistical

7–100% signal bin-by-bin 4–6%

uncertainties

Theory uncertainties in the signal acceptance

µrand µfvariations 1% signal norm. 1%

PDF 1% signal norm. 1%

Effect of b quark loop

3% signal norm. 3%

(17)

7.2 Uncertainties related to signal 15

in the standard model (19.3 pb). Under these assumptions, the contamination of region A by the signal is estimated to be less than 2% for all mass hypotheses mφ1 and in all bins of the

two-dimensional (m1, m2) distribution, with the exception of bin (4,4), where the contamination can

reach 12% for mφ1 = 8 GeV. However, the overall effect on the signal extraction is found to be

marginal. Within this conservative scenario, variations of C(i, j)due to possible contamination of control region A by the signal modify the observed and expected upper limits at 95% CL on

(σB)sigby less than 1% for all considered values of mφ1.

7.2 Uncertainties related to signal

The following uncertainties in the signal estimate are taken into account, and are summarized in Table 2.

An uncertainty of 2.6% is assigned to the integrated luminosity estimate [64].

The uncertainty in the muon identification and trigger efficiency is estimated to be 2% using the tag-and-probe technique applied to a sample of Z→µµ decays. Because final states with two

muons are selected in this analysis, this uncertainty translates into a 4% systematic uncertainty in the signal acceptance.

The track selection and isolation efficiency is assessed with a study performed on a sample of Z bosons decaying into a pair of τ leptons. In the selected Zττ events, one τ lepton

is identified via its muonic decay, while the other is identified as an isolated track resulting from a one-prong decay. The track is required to pass the nominal selection criteria used in the main analysis. From this study the uncertainty in the track selection and isolation efficiency is estimated to be 5%. As the analysis requires each muon to be accompanied by one track, this uncertainty gives rise to a 10% systematic uncertainty in the signal acceptance.

The muon momentum and track momentum scale uncertainties are smaller than 0.5% and have a negligible effect on the analysis.

The bin-by-bin MC statistical uncertainties in the signal acceptance range from 7 to 100%. Their impact on the signal normalization is between 4 and 6% as indicated in Table 1. These uncer-tainties are accounted for in the signal extraction procedure by 10 nuisance parameters, corre-sponding to 10 independent bins in the (m1, m2) distribution.

Theoretical uncertainties have an impact on the differential kinematic distributions of the pro-duced H(125) boson, in particular its pT spectrum, thereby affecting signal acceptance. The

uncertainty due to missing higher-order corrections to the gluon-gluon fusion process are esti-mated with the HQT program by varying the renormalization (µr) and factorization (µf) scales.

The H(125) pT-dependent k factors are recomputed according to these variations and applied

to the simulated signal samples. The resulting effect on the signal acceptance is estimated to be of the order of 1%.

The HQT program is also used to evaluate the effect of the PDF uncertainties. The nominal k factors for the H(125)boson pTspectrum are computed with the MSTW2008nnlo PDF set [65].

Variations of the MSTW2008nnlo PDFs within their uncertainties change the signal acceptance by about 1%, whilst using the CTEQ6L1 PDF set changes the signal acceptance by about 0.7%. These variations are covered by the assigned uncertainty of 1%.

The contribution of b quark loops to the gluon-gluon fusion process depends on the NMSSM parameters, in particular tan β, the ratio of the vacuum expectation values of the two NMSSM Higgs doublets. The corresponding uncertainty is conservatively estimated by calculating k factors for the H(125)boson pT spectrum with POWHEG [66–69], removing any contribution

(18)

16 8 Results

from the top quark loop and retaining only the contribution from the b quark loop. The modi-fied k factors applied to the simulated signal samples change the signal acceptance by approx-imately 3% for all mass hypotheses mφ1.

8

Results

The signal is extracted with a binned maximum-likelihood fit applied to the two-dimensional (m1, m2) distribution in data. For each mass hypothesis of the φ1boson, the (m1, m2) distribution

in data is fitted with the QCD multijet background shape and the gg→H(125)signal shape for the φ1mass under test. The contribution to the final selected sample from vector boson

fu-sion and vector boson associated production of the H(125)boson is suppressed by the selection described in Section 4, i.e. by the requirement∆R(µ, µ) >2. The impact of other backgrounds

on the fit is found to be negligible. The signal shapes are derived from simulation. The back-ground shape is evaluated from data, as described in Section 6. The systematic uncertainties are accounted for in the fit via nuisance parameters with log-normal pdfs.

The contribution to the final selected sample from vector boson fusion (qqH) and vector boson associated production (VH) of the H(125) boson is suppressed by the selection described in Section 4, especially by the requirement ∆R(µ, µ) > 2. For the values of the H(125) boson

production cross sections predicted in the SM, the expected contribution from the qqH and VH processes to the final selected sample is estimated to be less than 4% of total signal yield for all tested mφ1 hypotheses. The shapes of the two-dimensional (m1, m2) distributions are

found to be nearly indistinguishable among the three considered production modes, making it difficult to extract individual contributions from these processes in a model independent way. In the following these contributions are neglected, resulting in more conservative upper limits on (σB)sig. Subtraction of the qqH and VH contributions assuming the SM cross sections for

the H(125)production mechanisms would decrease the upper limits on(σB)sig by less than

4% for all tested values of mφ1.

First, the data are examined for their consistency with the background-only hypothesis by means of a fit performed with the normalization of the signal fixed to zero. Figure 7 (left) shows the two-dimensional (m1, m2) distribution unrolled into a one-dimensional array of

anal-ysis bins after performing the maximum-likelihood fit under the background-only hypothesis. The signal distribution, although not used in the fit, is also included for comparison, for the mass hypotheses mφ1 =4 and 8 GeV.

Table 3 presents the number of observed data events, the predicted background yields obtained from a fit under the background-only hypothesis, and the expected signal yields obtained from simulation, for each unique bin in the two-dimensional (m1, m2) distribution. The data are well

described by the background-only model.

The signal cross section times branching fraction is constrained by performing a fit under the signal+background hypothesis, where both the background and signal normalisations are al-lowed to vary freely in the fit. A representative example of the fit under the signal+background hypothesis at mφ1 = 8 GeV is presented in Fig. 7 (right). No significant deviations from the

background expectation are observed in data. Only a small excess is found for 6 ≤ mφ1

8 GeV, with a local significance ranging between 1.2σ (mφ1 = 8 GeV) and 1.4σ (mφ1 = 6 GeV). Results of the analysis are used to set upper limits on(σB)sig at 95% CL. The modified

fre-quentist CLs criterion [70, 71], implemented in the ROOSTATS package [72], is used for the

calculation of the exclusion limits. Figure 8 shows the observed upper limit on(σB)sigat 95%

(19)

17 Events / bin 0 50 100 150 200 250 300 (8 TeV) -1 19.7 fb CMS Data Bkg(+uncertainty) ) = 4 GeV 1 φ Signal, m( ) = 8 GeV 1 φ Signal, m( Bins (1,1) (1,2) (1,3) (1,4) (2,2) (2,3) (2,4) (3,3) (3,4) (4,4) Obs/Bkg 0.5 1 1.5 Events / bin 0 50 100 150 200 250 300 (8 TeV) -1 19.7 fb CMS Data Bkg(+uncertainty) ) = 8 GeV ) 1 φ Bkg+Signal ( m( Bins (1,1) (1,2) (1,3) (1,4) (2,2) (2,3) (2,4) (3,3) (3,4) (4,4) Obs/Bkg 0.5 1 1.5

Figure 7: The two-dimensional (m1, m2) distribution unrolled into a one-dimensional array of

analysis bins. In the left plot, data (points) are compared with the background prediction (solid histogram) after applying the maximum-likelihood fit under the background-only hypothe-sis and with the signal expectation for two mass hypotheses, mφ1 = 4 and 8 GeV (dotted and

dashed histograms, respectively). The signal distributions are obtained from simulation and normalized to a value of the cross section times branching fraction of 5 pb. In the right plot, data (points) are compared with the background prediction (solid histogram) and the back-ground+signal prediction for mφ1 = 8 GeV (dashed histogram) after applying the maximum-likelihood fit under the signal+background hypothesis. The bin notation follows the definition presented in Fig. 2.

(20)

18 9 Summary

Table 3: The number of observed data events, the predicted background yields, and the ex-pected signal yields, for different masses of the φ1boson in individual bins of the (m1, m2)

dis-tribution. The background yields and uncertainties are obtained from the maximum-likelihood fit under the background-only hypothesis. The signal yields are obtained from simulation and normalized to a signal cross section times branching fraction of 5 pb. The uncertainties in the signal yields include systematic and MC statistical uncertainties. The bin notation follows the definition presented in Fig. 2.

Bin Data Bkg. Signal for(σB)sig= 5 pb, mφ1 =

4 GeV 5 GeV 6 GeV 7 GeV 8 GeV

(1,1) 124 116±7 9.7±1.5 1.9±0.5 <0.1 0.1±0.1 <0.1 (1,2) 231 247±10 21.6±2.9 6.8±1.1 1.9±0.5 0.3±0.2 0.1±0.1 (1,3) 91 98±6 3.8±0.8 4.9±0.9 2.4±0.6 0.9±0.3 0.2±0.2 (1,4) 64 60±5 0.1±0.1 1.5±0.4 1.8±0.5 0.8±0.3 0.5±0.2 (2,2) 137 142±8 14.2±2.0 8.2±1.3 2.8±0.6 1.5±0.4 0.8±0.3 (2,3) 112 104±6 3.7±0.7 10.4±1.6 9.2±1.4 4.4±0.8 2.3±0.6 (2,4) 61 59±5 <0.1 2.6±0.6 5.6±1.0 8.1±1.3 4.0±0.8 (3,3) 16 19±2 <0.1 4.8±0.9 4.8±0.9 3.7±0.7 2.2±0.5 (3,4) 29 23±3 <0.1 1.9±0.5 8.0±0.9 11.1±1.5 9.4±1.4 (4,4) 8 7±1 <0.1 <0.1 3.1±0.6 9.1±1.4 11.2±1.7

Table 4: The observed upper limit on(σB)sig at 95% CL, together with the expected limit

ob-tained in the background-only hypothesis, as a function of mφ1. Also shown are±1σ and±

probability intervals around the expected limit. mφ1[GeV]

Upper limits on(σB)sig[pb] at 95% CL

observed − expected + + 4 7.1 5.7 7.6 10.6 14.9 20.2 5 10.3 5.4 7.3 10.3 15.0 21.2 6 8.6 2.8 3.8 5.4 7.8 11.0 7 5.0 1.6 2.2 3.1 4.5 6.5 8 4.5 1.5 2.0 2.9 4.3 6.2

in the range from 4 to 8 GeV. Exclusion limits are also reported in Table 4.

The observed limit is compatible with the expected limit within two standard deviations in the entire tested range of the φ1 boson mass, 4 ≤ mφ1 ≤ 8 GeV. The observed limit ranges from

4.5 pb at mφ1 = 8 GeV to 10.3 pb at mφ1 = 5 GeV. The expected limit ranges from 2.9 pb at

mφ1 =8 GeV to 10.6 pb at mφ1 =4 GeV.

The analysis presented here complements the search for h/H → a1a1 → µµττperformed by

the ATLAS Collaboration [52], providing results in the 4τ channel, which has not been previ-ously explored at the LHC.

9

Summary

A search for a very light NMSSM Higgs boson a1 or h1, produced in decays of the observed

boson with a mass near 125 GeV, H(125), is performed on a pp collision data set corresponding to an integrated luminosity of 19.7 fb−1, collected at a centre-of-mass energy of 8 TeV. The anal-ysis searches for the production of an H(125)boson via gluon-gluon fusion, and its decay into

(21)

19

(GeV)

1 φ

m

4

5

6

7

8

(pb)

sig

)

B

σ

U

p

p

e

r

9

5

%

C

L

l

im

it

o

n

(

0

2

4

6

8

10

12

14

16

18

20

Observed Expected Expected σ 1 ± Expected σ 2 ± (8 TeV) -1 19.7 fb

CMS

Observed Expected Expected σ 1 ± Expected σ 2 ±

Figure 8: The observed and expected upper limits on(σB)sig in pb at 95% CL, as a function of

1. The expected limit is obtained under the background-only hypothesis. The bands show the expected±1σ and±2σ probability intervals around the expected limit.

a pair of a1(h1)states, each of which decays into a pair of τ leptons. The search covers a mass

range of the a1 (h1)boson of 4 to 8 GeV. No significant excess above background expectations

is found in data, and upper limits at 95% CL are set on the signal production cross section times branching fraction,

(σB)sig≡σ(gg→H(125)) B(H(125) →φ1φ1) B2(φ1 →ττ),

where φ1is either the a1 or h1boson. The observed upper limit at 95% CL on (σB)sig ranges

from 4.5 pb at mφ1 =8 GeV to 10.3 pb at mφ1 =5 GeV.

Acknowledgements

We congratulate our colleagues in the CERN accelerator departments for the excellent perfor-mance of the LHC and thank the technical and administrative staffs at CERN and at other CMS institutes for their contributions to the success of the CMS effort. In addition, we grate-fully acknowledge the computing centres and personnel of the Worldwide LHC Computing Grid for delivering so effectively the computing infrastructure essential to our analyses. Fi-nally, we acknowledge the enduring support for the construction and operation of the LHC and the CMS detector provided by the following funding agencies: the Austrian Federal Min-istry of Science, Research and Economy and the Austrian Science Fund; the Belgian Fonds de la Recherche Scientifique, and Fonds voor Wetenschappelijk Onderzoek; the Brazilian Fund-ing Agencies (CNPq, CAPES, FAPERJ, and FAPESP); the Bulgarian Ministry of Education and Science; CERN; the Chinese Academy of Sciences, Ministry of Science and Technology, and Na-tional Natural Science Foundation of China; the Colombian Funding Agency (COLCIENCIAS);

(22)

20 References

the Croatian Ministry of Science, Education and Sport, and the Croatian Science Foundation; the Research Promotion Foundation, Cyprus; the Ministry of Education and Research, Esto-nian Research Council via IUT23-4 and IUT23-6 and European Regional Development Fund, Estonia; the Academy of Finland, Finnish Ministry of Education and Culture, and Helsinki Institute of Physics; the Institut National de Physique Nucl´eaire et de Physique des Partic-ules / CNRS, and Commissariat `a l’ ´Energie Atomique et aux ´Energies Alternatives / CEA, France; the Bundesministerium f ¨ur Bildung und Forschung, Deutsche Forschungsgemeinschaft, and Helmholtz-Gemeinschaft Deutscher Forschungszentren, Germany; the General Secretariat for Research and Technology, Greece; the National Scientific Research Foundation, and Na-tional Innovation Office, Hungary; the Department of Atomic Energy and the Department of Science and Technology, India; the Institute for Studies in Theoretical Physics and Mathe-matics, Iran; the Science Foundation, Ireland; the Istituto Nazionale di Fisica Nucleare, Italy; the Ministry of Science, ICT and Future Planning, and National Research Foundation (NRF), Republic of Korea; the Lithuanian Academy of Sciences; the Ministry of Education, and Uni-versity of Malaya (Malaysia); the Mexican Funding Agencies (CINVESTAV, CONACYT, SEP, and UASLP-FAI); the Ministry of Business, Innovation and Employment, New Zealand; the Pakistan Atomic Energy Commission; the Ministry of Science and Higher Education and the National Science Centre, Poland; the Fundac¸˜ao para a Ciˆencia e a Tecnologia, Portugal; JINR, Dubna; the Ministry of Education and Science of the Russian Federation, the Federal Agency of Atomic Energy of the Russian Federation, Russian Academy of Sciences, and the Russian Foun-dation for Basic Research; the Ministry of Education, Science and Technological Development of Serbia; the Secretar´ıa de Estado de Investigaci ´on, Desarrollo e Innovaci ´on and Programa Consolider-Ingenio 2010, Spain; the Swiss Funding Agencies (ETH Board, ETH Zurich, PSI, SNF, UniZH, Canton Zurich, and SER); the Ministry of Science and Technology, Taipei; the Thailand Center of Excellence in Physics, the Institute for the Promotion of Teaching Science and Technology of Thailand, Special Task Force for Activating Research and the National Sci-ence and Technology Development Agency of Thailand; the Scientific and Technical Research Council of Turkey, and Turkish Atomic Energy Authority; the National Academy of Sciences of Ukraine, and State Fund for Fundamental Researches, Ukraine; the Science and Technology Facilities Council, UK; the US Department of Energy, and the US National Science Foundation. Individuals have received support from the Marie-Curie programme and the European Re-search Council and EPLANET (European Union); the Leventis Foundation; the A. P. Sloan Foundation; the Alexander von Humboldt Foundation; the Belgian Federal Science Policy Of-fice; the Fonds pour la Formation `a la Recherche dans l’Industrie et dans l’Agriculture (FRIA-Belgium); the Agentschap voor Innovatie door Wetenschap en Technologie (IWT-(FRIA-Belgium); the Ministry of Education, Youth and Sports (MEYS) of the Czech Republic; the Council of Science and Industrial Research, India; the HOMING PLUS programme of the Foundation for Polish Science, cofinanced from European Union, Regional Development Fund; the OPUS programme of the National Science Center (Poland); the Compagnia di San Paolo (Torino); the Consorzio per la Fisica (Trieste); MIUR project 20108T4XTM (Italy); the Thalis and Aristeia programmes cofinanced by EU-ESF and the Greek NSRF; the National Priorities Research Program by Qatar National Research Fund; the Rachadapisek Sompot Fund for Postdoctoral Fellowship, Chula-longkorn University (Thailand); and the Welch Foundation, contract C-1845.

References

[1] ATLAS Collaboration, “Observation of a new particle in the search for the Standard Model Higgs boson with the ATLAS detector at the LHC”, Phys. Lett. B 716 (2012) 1, doi:10.1016/j.physletb.2012.08.020, arXiv:1207.7214.

(23)

References 21

[2] CMS Collaboration, “Observation of a new boson at a mass of 125 GeV with the CMS experiment at the LHC”, Phys. Lett. B 716 (2012) 30,

doi:10.1016/j.physletb.2012.08.021, arXiv:1207.7235.

[3] CMS Collaboration, “Observation of a new boson with mass near 125 GeV in pp collisions at√s = 7 and 8 TeV”, JHEP 06 (2013) 081,

doi:10.1007/JHEP06(2013)081, arXiv:1303.4571.

[4] CMS Collaboration, “Precise determination of the mass of the Higgs boson and tests of compatibility of its couplings with the standard model predictions using proton collisions at 7 and 8 TeV”, Eur. Phys. J. C 75 (2015) 212,

doi:10.1140/epjc/s10052-015-3351-7, arXiv:1412.8662.

[5] CMS Collaboration, “Observation of the diphoton decay of the Higgs boson and measurement of its properties”, Eur. Phys. J. C 74 (2014) 3076,

doi:10.1140/epjc/s10052-014-3076-z, arXiv:1407.0558.

[6] CMS Collaboration, “Measurement of the properties of a Higgs boson in the four-lepton final state”, Phys. Rev. D 89 (2014) 092007, doi:10.1103/PhysRevD.89.092007, arXiv:1312.5353.

[7] CMS Collaboration, “Measurement of Higgs boson production and properties in the WW decay channel with leptonic final states”, JHEP 01 (2014) 096,

doi:10.1007/JHEP01(2014)096, arXiv:1312.1129.

[8] CMS Collaboration, “Evidence for the 125 GeV Higgs boson decaying to a pair of

τleptons”, JHEP 05 (2014) 104, doi:10.1007/JHEP05(2014)104,

arXiv:1401.5041.

[9] CMS Collaboration, “Evidence for the direct decay of the 125 GeV Higgs boson to

fermions”, Nature Phys. 10 (2014) 557, doi:10.1038/nphys3005, arXiv:1401.6527. [10] ATLAS Collaboration, “Measurement of Higgs boson production in the diphoton decay

channel in pp collisions at center-of-mass energies of 7 and 8 TeV with the ATLAS detector”, Phys. Rev. D 90 (2014) 112015, doi:10.1103/PhysRevD.90.112015, arXiv:1408.7084.

[11] ATLAS Collaboration, “Measurements of Higgs boson production and couplings in the four-lepton channel in pp collisions at center-of-mass energies of 7 and 8 TeV with the ATLAS detector”, Phys. Rev. D 91 (2015) 012006,

doi:10.1103/PhysRevD.91.012006, arXiv:1408.5191.

[12] ATLAS Collaboration, “Observation and measurement of Higgs boson decays to WW∗ with the ATLAS detector”, Phys. Rev. D 92 (2015) 012006,

doi:10.1103/PhysRevD.92.012006, arXiv:1412.2641.

[13] ATLAS Collaboration, “Evidence for the Higgs-boson Yukawa coupling to tau leptons with the ATLAS detector”, JHEP 04 (2015) 117, doi:10.1007/JHEP04(2015)117, arXiv:1501.04943.

[14] ATLAS Collaboration, “Measurement of the Higgs boson mass from the H →γγand

H→ ZZ∗ →4`channels in pp collisions at center-of-mass energies of 7 and 8 TeV with the ATLAS detector”, Phys. Rev. D 90 (2014) 052004,

(24)

22 References

[15] Y. A. Gol’fand and E. P. Likhtman, “Extension of the algebra of Poincar´e group generators and violation of P invariance”, JETP Lett. 13 (1971) 323.

[16] J. Wess and B. Zumino, “Supergauge transformations in four dimensions”, Nucl. Phys. B

70(1974) 39, doi:10.1016/0550-3213(74)90355-1.

[17] J. R. Ellis, “Limits of the standard model”, (2002). arXiv:hep-ph/0211168. Lectures given at the PSI Summer School, Zuoz.

[18] P. Fayet, “Supergauge invariant extension of the Higgs mechanism and a model for the electron and its neutrino”, Nucl. Phys. B 90 (1975) 104,

doi:10.1016/0550-3213(75)90636-7.

[19] P. Fayet, “Spontaneously broken supersymmetric theories of weak, electromagnetic and strong interactions”, Phys. Lett. B 69 (1977) 489,

doi:10.1016/0370-2693(77)90852-8.

[20] R. K. Kaul and P. Majumdar, “Cancellation of quadratically divergent mass corrections in globally supersymmetric spontaneously broken gauge theories”, Nucl. Phys. B 199 (1982) 36, doi:10.1016/0550-3213(82)90565-X.

[21] R. Barbieri, S. Ferrara, and C. A. Savoy, “Gauge models with spontaneously broken local supersymmetry”, Phys. Lett. B 119 (1982) 343,

doi:10.1016/0370-2693(82)90685-2.

[22] H. P. Nilles, M. Srednicki, and D. Wyler, “Weak interaction breakdown induced by supergravity”, Phys. Lett. B 120 (1983) 346, doi:10.1016/0370-2693(83)90460-4. [23] J.-M. Frere, D. R. T. Jones, and S. Raby, “Fermion masses and induction of the weak scale

by supergravity”, Nucl. Phys. B 222 (1983) 11, doi:10.1016/0550-3213(83)90606-5.

[24] J.-P. Derendinger and C. A. Savoy, “Quantum effects and SU(2)×U(1) breaking in supergravity gauge theories”, Nucl. Phys. B 237 (1984) 307,

doi:10.1016/0550-3213(84)90162-7.

[25] U. Ellwanger, C. Hugonie, and A. M. Teixeira, “The Next-to-Minimal Supersymmetric Standard Model”, Phys. Rept. 496 (2010) 1, doi:10.1016/j.physrep.2010.07.001, arXiv:0910.1785.

[26] M. Maniatis, “The Next-to-Minimal Supersymmetric extension of the Standard Model reviewed”, Int. J. Mod. Phys. A 25 (2010) 3505, doi:10.1142/S0217751X10049827, arXiv:0906.0777.

[27] J. E. Kim and H. P. Nilles, “The µ-problem and the strong CP-problem”, Phys. Lett. B 138 (1984) 150, doi:10.1016/0370-2693(84)91890-2.

[28] J. A. Casas, J. R. Espinosa, and I. Hidalgo, “The MSSM fine tuning problem: a way out”, JHEP 01 (2004) 008, doi:10.1088/1126-6708/2004/01/008,

arXiv:hep-ph/0310137.

[29] R. Dermisek and J. F. Gunion, “Escaping the Large Fine-Tuning and Little Hierarchy Problems in the Next to Minimal Supersymmetric Model and h→aa Decays”, Phys. Rev. Lett. 95 (2005) 041801, doi:10.1103/PhysRevLett.95.041801,

(25)

References 23

[30] R. Dermisek and J. F. Gunion, “Next-to-minimal supersymmetric model solution to the fine-tuning problem, precision electroweak constraints, and the largest CERN LEP Higgs event excess”, Phys. Rev. D 76 (2007) 095006, doi:10.1103/PhysRevD.76.095006, arXiv:0705.4387.

[31] G. Belanger et al., “Higgs Bosons at 98 and 125 GeV at LEP and the LHC”, JHEP 01 (2013) 069, doi:10.1007/JHEP01(2013)069, arXiv:1210.1976.

[32] G. Belanger et al., “Two Higgs Bosons at the Tevatron and the LHC?”, (2012). arXiv:1208.4952.

[33] J. F. Gunion, Y. Jiang, and S. Kraml, “Diagnosing Degenerate Higgs Bosons at 125 GeV”, Phys. Rev. Lett. 110 (2013) 051801, doi:10.1103/PhysRevLett.110.051801, arXiv:1208.1817.

[34] J. F. Gunion, Y. Jiang, and S. Kraml, “Could two NMSSM Higgs bosons be present near 125 GeV?”, Phys. Rev. D 86 (2012) 071702, doi:10.1103/PhysRevD.86.071702, arXiv:1207.1545.

[35] S. F. King, M. M ¨uhlleitner, and R. Nevzorov, “NMSSM Higgs benchmarks near 125 GeV”, Nucl. Phys. B 860 (2012) 207, doi:10.1016/j.nuclphysb.2012.02.010,

arXiv:1201.2671.

[36] S. F. King, M. M ¨uhlleitner, R. Nevzorov, and K. Walz, “Natural NMSSM Higgs bosons”, Nucl. Phys. B 870 (2013) 323, doi:10.1016/j.nuclphysb.2013.01.020,

arXiv:1211.5074.

[37] D. Curtin et al., “Exotic decays of the 125 GeV Higgs boson”, Phys. Rev. D 90 (2014) 075004, doi:10.1103/PhysRevD.90.075004, arXiv:1312.4992.

[38] U. Ellwanger, J. F. Gunion, and C. Hugonie, “Difficult scenarios for NMSSM Higgs discovery at the LHC”, JHEP 07 (2005) 041,

doi:10.1088/1126-6708/2005/07/041, arXiv:hep-ph/0503203.

[39] U. Ellwanger, J. F. Gunion, C. Hugonie, and S. Moretti, “Towards a No-Lose Theorem for NMSSM Higgs Discovery at the LHC”, in Physics at TeV colliders, Les Houches workshop. 2003. arXiv:hep-ph/0305109.

[40] U. Ellwanger, J. F. Gunion, C. Hugonie, and S. Moretti, “NMSSM Higgs Discovery at the LHC”, in Physics at TeV colliders, Les Houches workshop. 2003. arXiv:hep-ph/0401228. [41] A. Belyaev et al., “The Scope of the 4 tau Channel in Higgs-strahlung and Vector Boson

Fusion for the NMSSM No-Lose Theorem at the LHC”, (2008). arXiv:0805.3505. [42] A. Belyaev et al., “LHC discovery potential of the lightest NMSSM Higgs boson in the

h1 →a1a1→4µ channel”, Phys. Rev. D 81 (2010) 075021,

doi:10.1103/PhysRevD.81.075021, arXiv:1002.1956.

[43] M. Lisanti and J. G. Wacker, “Discovering the Higgs boson with low mass muon pairs”, Phys. Rev. D 79 (2009) 115006, doi:10.1103/PhysRevD.79.115006,

arXiv:0903.1377.

[44] M. M. Almarashi and S. Moretti, “Scope of Higgs production in association with a bottom quark pair in probing the Higgs sector of the NMSSM at he LHC”, (2012). arXiv:1205.1683.

Imagem

Figure 1: Left: Feynman diagram for the signal process. Right: Illustration of the signal topol- topol-ogy
Table 1: The number of observed events, expected background and signal yields, and signal ac- ac-ceptances after final selection
Figure 2: Binning of the two-dimensional (m 1 , m 2 ) distribution. The hatched bins are excluded from the statistical analysis, as detailed in the text.
Figure 3: Normalized invariant mass distributions of the first muon and the softest (left plots) or hardest (right plots) accompanying track for different isolation requirements imposed on the second muon: when the second muon has only one accompanying tra
+7

Referências

Documentos relacionados

Atraves do estudo realizado, constata-se com relacao as praticas mais importantes que identificam uma empresa socialmente responsavel, que 64,5% dos estudantes identificam uma

Additionally, the chemical composition of fig peel justifies the recovery of this part of the fruit for further application in food industry, in the cases where it is discarded for

Eles se explicam somen­ te pela paixão dos seus autores, a qual aparece sob as a parências de ob­ jetividade e de seren id

Gnatópodo 2 com palma transversa definida por um grande dente seguido por uma escavação profunda e mais dois dentes, artículo 7 com projeção arredondada na margem

19th EANA Astrobiology Conference – Orléans, France The European Astrobiology Network Association, EANA, joins together people insterested in the origins of life and

No entanto, foi notório que o anseio pela mudança na portaria dos condomínios parte inicialmente de experiências e fatores externos ao da solução, como a

Determinacion de la entrada en pubertad en cabras Blancas Celiiber i cas naci d as en Olano media n te. observacion de celos y anal isis de

A umidade presente no ar pode promover o reinício das atividades do embrião, caso o oxigênio e a temperatura sejam suficientes para que tal processo aconteça, acelerando, dessa